Representational similarity analysis connecting the branches of systems neuroscience

Size: px
Start display at page:

Download "Representational similarity analysis connecting the branches of systems neuroscience"

Transcription

1 SYSTEMS NEUROSCIENCE ORIGINAL RESEARCH ARTICLE published: 24 November 2008 doi: /neuro connecting the branches of systems neuroscience Nikolaus Kriegeskorte 1, *, Marieke Mur 1,2 and Peter Bandettini 1 1 Section on Functional Imaging Methods, Laboratory of Brain and Cognition, National Institute of Mental Health, National Institutes of Health, Bethesda, MD, USA 2 Department of Cognitive Neuroscience, Faculty of Psychology, Maastricht University, Maastricht, The Netherlands Edited by: Mriganka Sur, Massachusetts Institute of Technology, USA Reviewed by: Michael A. Silver, University of California, USA Doris Y. Tsao, University of Bremen, Germany *Correspondence: Nikolaus Kriegeskorte, Section on Functional Imaging Methods, Laboratory of Brain and Cognition, National Institute of Mental Health, National Institutes of Health, Building 10, Room 1D80B, 10 Center Dr. MSC 1148, Bethesda, MD , USA. kriegeskorten@mail.nih.gov A fundamental challenge for systems neuroscience is to quantitatively relate its three major branches of research: brain-activity measurement, behavioral measurement, and computational modeling. Using measured brain-activity patterns to evaluate computational network models is complicated by the need to define the correspondency between the units of the model and the channels of the brain-activity data, e.g., single-cell recordings or voxels from functional magnetic resonance imaging (fmri). Similar correspondency problems complicate relating activity patterns between different modalities of brain-activity measurement (e.g., fmri and invasive or scalp electrophysiology), and between subjects and species. In order to bridge these divides, we suggest abstracting from the activity patterns themselves and computing representational dissimilarity matrices (RDMs), which characterize the information carried by a given representation in a brain or model. Building on a rich psychological and mathematical literature on similarity analysis, we propose a new experimental and data-analytical framework called representational similarity analysis (RSA), in which multi-channel measures of neural activity are quantitatively related to each other and to computational theory and behavior by comparing RDMs. We demonstrate RSA by relating representations of visual objects as measured with fmri in early visual cortex and the fusiform face area to computational models spanning a wide range of complexities. The RDMs are simultaneously related via second-level application of multidimensional scaling and tested using randomization and bootstrap techniques. We discuss the broad potential of RSA, including novel approaches to experimental design, and argue that these ideas, which have deep roots in psychology and neuroscience, will allow the integrated quantitative analysis of data from all three branches, thus contributing to a more unified systems neuroscience. Keywords: fmri, electrophysiology, computational modeling, population code, similarity, representation INTRODUCTION RELATING REPRESENTATIONS IN BRAINS AND MODELS A computational model of a single neuron (e.g., in V1) can be tested and adjusted on the basis of electrophysiological recordings of the activity of that type of neuron under a variety of circumstances (e.g., across different stimuli). This has been one successful avenue of evaluating computational models of single neurons with brain-activity data (e.g., David and Gallant, 2005; Koch, 1999; Rieke et al., 1999). This single-unit fitting approach becomes intractable, however, for computational models at a larger scale of organization, which simulate comprehensive brain information processing and include populations of units with different functional properties. A major problem in relating such models to brain-activity data is the spatial correspondency problem: Which single-cell recording or functional magnetic resonance imaging (fmri) voxel corresponds to which unit of the computational model? Defining a one-to-one mapping between model units and data channels will require that the functional properties of the simulated and real neurons are well characterized in advance; and finding the optimal match-up will still be challenging. To further complicate matters, a one-to-one mapping often cannot be assumed in the first place; the voxels and sensors of brain imaging, for example, reflect the activity of large numbers of neurons. Although model units as well can represent sets of neurons, we cannot in general assume a one-to-one correspondency. When a one-to-one mapping does not exist, the attempt to define such a mapping is clearly ill-motivated. Defining the correspondency more generally in terms of a linear transform would require the fitting of a weights matrix, which will often have a prohibitively large number of parameters (number of model units by number of data channels). Similar correspondency problems arise in relating activity patterns between different modalities of brain-activity measurement. Modern techniques of multi-channel brain-activity measurement (including invasive and scalp electrophysiology, as well as fmri) can take rich samples of neuronal pattern information. Invasive electrophysiology is the ideal modality in terms of resolution in both space (single neuron) and time (ms). However, only a very small subset of neurons can be recorded from simultaneously. Imaging techniques (fmri and scalp electrophysiology), sample neuronal activity contiguously across large parts of the brain or across the Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 1

2 whole brain. In imaging, however, a single channel reflects the joint activity of tens of thousands (high-resolution fmri), or even millions of neurons (scalp electrophysiology). If the same activity patterns are measured with two different techniques, we expect an overlap in the information sampled. However, different techniques sample activity patterns in fundamentally different ways. Invasive electrophysiology measures the electrical activity of single cells, whereas fmri measures the hemodynamic aspect of brain activity. Although the hemodynamic fmri signal has been shown to reflect neuronal activity (Logothetis et al. 2001; see also Bandettini and Ungerleider, 2001), fmri patterns are spatiotemporally displaced, smoothed, and distorted. Scalp electrophysiology combines high temporal resolution with a spatial sampling of neuronal activity that is even coarser than in fmri. Neuroscientific theory must abstract from the idiosyncrasies of particular empirical modalities. To this end, we need a modalityindependent way of characterizing a brain region s representation. Such a characterization will also enable us to elucidate in how far different modalities provide consistent or inconsistent information. One way of characterizing the information a brain region represents is in terms of the mental states (e.g., stimulus percepts) it distinguishes (Figure 1). Here we suggest to relate modalities of brain-activity measurement and information-processing models by comparing activity-pattern dissimilarity matrices. Our approach FIGURE 1 Characterizing brain regions by representational similarity structure. For each region, a similarity-graph icon shows the similarities between the activity patterns elicited by four stimulus images. Images placed close together in the icon elicited similar response patterns. Images placed far apart elicited dissimilar response patterns. The color of each connection line indicates whether the response-pattern difference was significant for the group (red: p < 0.01; light gray: p 0.05, not significant). A connection line, like a rubberband, becomes thinner when stretched beyond the length that would exactly reflect the dissimilarity it represents. Connections also become thicker when compressed. Line thickness, thus, indicates the inevitable distortion of the 2D representation of the higher-dimensional similarity structure. The thickness of the connection lines is chosen such that the area of each connection (length times thickness) precisely reflects the dissimilarity measure. This novel visualization of fmri response-pattern information combines (A) a multidimensional-scaling arrangement of activity-pattern similarity (as introduced to fmri by Edelman et al., 1998), (B) a novel rubberband-graph depiction of inevitable distortions, and (C) the results of statistical tests of a patterninformation analysis (for details on the test, see Kriegeskorte et al., 2007). The icons show fixed-effects group analyses for regions of interest individually defined in 11 subjects. Early visual cortex was anatomically defined; all other regions were functionally defined using a data set independent of that used to compute the similarity-graph icons and statistical tests. Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 2

3 FIGURE 2 Computation of the representational dissimilarity matrix. For each pair of experimental conditions, the associated activity patterns (in a brain region or model) are compared by spatial correlation. The dissimilarity between them is measured as 1 minus the correlation (0 for perfect correlation, 1 for no correlation, 2 for perfect anticorrelation). These dissimilarities for all pairs of conditions are assembled in the RDM. Each cell of the RDM, thus, compares the response patterns elicited by two images. As a consequence, an RDM is symmetric about a diagonal of zeros. To visualize the representation for a small number of conditions, we suggest the similarity-graph icon (top right, cf. Figure 1). obviates the need for defining explicit spatial correspondency mappings or transformations from one modality into another. THE REPRESENTATIONAL DISSIMILARITY MATRIX For a given brain region, we interpret (Dennett, 1987) the activity pattern associated with each experimental condition as a representation (e.g., a stimulus representation) 1. By comparing the activity patterns associated with each pair of conditions (Edelman 1 More generally, we can think of the activity pattern as the physical manifestation of the mental state induced by the experimental condition. The mental state could be the percept of an external object or something more remotely related to the external world, such as a motor program, a plan, or an emotion. et al., 1998; Haxby et al., 2001), we obtain a representational dissimilarity matrix (RDM; Figure 2), which serves to characterize the representation 2. An RDM contains a cell for each pair of experimental conditions (Figure 2). Each cell contains a number reflecting the dissimilarity between the activity patterns associated with the two conditions. As a consequence, an RDM is symmetric about a diagonal of 2 Note that similarity (a term we use here to refer to the general concept) can equally well be characterized by a similarity measure (in which greater values indicate greater similarity) or a dissimilarity measure (in which greater values indicate less similarity). We prefer the latter because of its intuitive relationship to distances in a multidimensional space. Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 3

4 zeros. We suggest using correlation distance (1-correlation) as the dissimilarity measure, although we explore a number of measures below (Figure 10). The RDM indicates the degree to which each pair of conditions is distinguished. It can thus be viewed as encapsulating the information content (in an informal sense) carried by the region. For any computational model (Figure 5) that can be exposed to the same experimental conditions (e.g., presented with the same stimuli), we can obtain an RDM for each of its processing stages in the same way as for a brain region (Figure 6). The RDMs serve as the signatures of regional representations in brains and models. Importantly, these signatures abstract from the spatial layout of the representations. They are indexed (horizontally and vertically) by experimental condition and can thus be directly compared between brain and model. What we are comparing, intuitively, is the represented information, not the activity patterns themselves. MATCHING DISSIMILARITY MATRICES: A SECOND-ORDER ISOMORPHISM RDMs can be quantitatively compared just like activity patterns, e.g., using correlation distance (1-correlation) or rank-correlation distance. Because RDMs are symmetric about a diagonal of zeros, we will apply these measures using only the upper (or equivalently the lower) triangle of the matrices. Analysis of similarity structure has a history in psychology and related fields. When exposed to a suitable sensory stimulus, our brain activity reflects many properties of the stimulus. The reflection of a stimulus property in the activity level of a neuron constitutes what has been termed a first-order isomorphism between the property and its representation in the brain. Most neuroscientific studies of brain representations have focused on the relationship between stimulus properties and brain-activity level in single cells or brain regions, i.e., on the first-order isomorphism between stimuli and their representations. One concept at the core of our approach is that of second-order isomorphism (Shepard and Chipman, 1970), i.e., the match of dissimilarity matrices. When we encounter difficulty establishing a direct correspondence, i.e., a first-order isomorphism 3, in studying the relationship between stimuli and their representations, we may attempt instead to establish a correspondence between the relations among the stimuli on the one hand and the relations among their representations on the other, i.e., a second-order isomorphism. We can study the second-order isomorphism by relating the similarity structure 3 A first-order isomorphism between object and representation can be interpreted in several ways. Naively: The representation is a replication of the object, i.e., identical with it. (Problem: A chair does not fit into the human skull.) More reasonably, we may interpret first-order isomorphism as a mere similarity of some sort. For example a retinotopic representation of an image in V1 may emit no light, be smaller and distorted, but it does bear a topological similarity to the image. More cautiously, we could maintain that first-order isomorphism requires only that the representation has properties (e.g., neuronal firing rates) that are related to properties of the objects represented (e.g., line orientation). While the naive interpretation is clearly untenable, the other interpretations are generally accepted in neuroscience. We concur with this widespread view, which motivates studies of stimulus selectivity at the level of single cells and brain regions. However, we feel that analysis of the second-order isomorphism (which can reflect a first-order isomorphism) is equally promising and offers a complementary higher-level functional perspective. of the objects to the similarity structure of the representations. This promises a higher-level functional perspective, which is complementary to the perspective of first-order isomorphism. RELATED APPROACHES IN THE LITERATURE The qualitative and quantitative analysis of similarity structure has a long history in philosophy, psychology, and neuroscience. A good entry to the literature is provided by Edelman (1998), who (Edelman et al., 1998) also pioneered application of similarity analysis to fmri activity patterns using the technique of multidimensional scaling (MDS; Borg and Groenen, 2005; Kruskal and Wish, 1978; Shepard, 1980; Torgerson, 1958). Laakso and Cottrell (2000) compared representations in hidden units of connectionist networks by correlating the dissimilarity structures of their activity patterns. They suggest that this approach could be used as a general method for comparing representations and discuss the philosophical implications. Op de Beeck et al. (2001) related the representational similarity of silhouette shapes in monkey inferior temporal cortex to physical and behavioral similarity measures for those stimuli. At a more general level, activity-pattern similarity is related to activity-pattern information as targeted in a number of recent studies in human fmri (Carlson et al., 2003; Cox and Savoy, 2003; Davatzikos et al., 2005; Friston et al., 2008; Hanson et al., 2004; Haxby et al., 2001; Haynes and Rees, 2005a,b; Haynes et al., 2007; Kamitani and Tong, 2005, 2006; Kriegeskorte et al., 2006; LaConte et al., 2005; Mitchell et al., 2004; Mourao-Miranda et al., 2005; Pessoa and Padmala, 2006; Polyn et al., 2005; Serences and Boynton, 2007; Spiridon and Kanwisher, 2002; Strother et al., 2002; Williams et al., 2007; for reviews see Haynes and Rees, 2006; Kriegeskorte and Bandettini, 2007; Norman et al., 2006) and also in monkey electrophysiology (Hung et al., 2005; Tsao et al., 2006). Explicit similarity analyses of neuronal activity patterns have begun to be applied in human fmri (Aguirre, 2007; Aguirre et al., in preparation; Drucker and Aguirre, submitted; Edelman et al., 1998; Kriegeskorte et al., in press; O Toole et al., 2005) and monkey electrophysiology (Kiani et al., 2007; Op de Beeck et al., 2001). CONNECTING THE BRANCHES OF SYSTEMS NEUROSCIENCE In this paper, we argue that the theoretical concept of second-order isomorphism (Shepard and Chipman, 1970) can serve a much more general purpose than previously thought, relating not only external objects to their brain representations, but bridging the divides between the three branches of systems neuroscience: behavioral experimentation, brain-activity experimentation, and computational modeling (Figure 3). We introduce an analysis framework called representational similarity analysis (RSA), which builds on a rich psychological and mathematical literature (Edelman, 1995, 1998; Edelman and Duvdevani-Bar, 1997a,b; Kruskal and Wish, 1978; Laakso and Cottrell, 2000; Shepard, 1980; Shepard and Chipman, 1970; Shepard et al., 1975; Torgerson, 1958). The core idea is to use the RDM as a signature of the representations in brain regions and computational models. We define a specific working prototype of RSA and discuss the potential of this approach in its full breadth: (1) Integration of computational modeling into the analysis of brain-activity data. A key advantage of RSA is that computational Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 4

5 models of brain information processing form an integrated component of data analysis and can be directly evaluated and compared. We demonstrate how to apply multivariate analysis to a set of dissimilarity matrices from brain regions and models in order to find out (1) which model best explains the representation in each brain region and (2) to what extent representations among regions and models resemble each other. We introduce a randomization test of representational relatedness and a bootstrap technique for obtaining error bars on estimates of the goodness of fit of different models. (2) Relating regions, subjects, species, and modalities of brainactivity measurement. We discuss how RSA can be used to quantitatively relate: representations in different regions of the same brain ( representational connectivity ), A B FIGURE 3 The representational dissimilarity matrix as a hub that relates different representations. (A) Systems neuroscience has struggled to relate its three major branches of research: behavioral experimentation, brain-activity experimentation, and computational modeling. So far these branches have interacted largely on two levels: (1) They have interacted on the level of verbal theory, i.e., by comparing conclusions drawn from separate analyses. This level is essential, but it is not quantitative. (2) They have interacted at the level characteristic functions, e.g., by comparing psychometric and neurometric functions. This form of bringing the branches in touch is equally essential and can be quantitative. However, characteristic functions typically contain only a small number of data points, so the interface is not informationally rich. Note that the RDM shown is based on only four conditions, yielding only (4 2 4)/2 = 6 parameters. However, since the number of parameters grows as the square of the number of conditions, the RDM can provide an informationally rich interface for relating different representations. Consider for example the 96-image experiment we discuss, where the matrix has ( )/2 = 4,560 parameters. (B) This panel illustrates in greater detail what different representations can be related via the quantitative interface provided by the RDM. We arbitrarily chose the example of fmri to illustrate the within-modality relationships that can be established. Note that all these relationships are difficult to establish otherwise (gray double arrows). Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 5

6 corresponding brain regions in different subjects ( intersubject information ), corresponding brain regions in different species (e.g., humans and monkeys), and different modalities of brain-activity data (e.g., cell recordings and fmri). (3) Relating brain and behavior. We discuss how RSA can quantitatively relate brain-activity measurements to behavioral data. This possibility has already been demonstrated in previous work (Aguirre et al., in preparation; Kiani et al., 2007; Op de Beeck et al., 2001). (4) Addressing a broader array of neuroscientific questions with each experiment by means of condition-rich design. While RSA is applicable to conventional experimental designs, is synergizes with novel condition-rich experimental designs, where a single experiment can address a large number of neuroscientific questions. We demonstrate this with an fmri experiment that has 96 separate conditions and discuss the broader implications. We hope that RSA will contribute to a more integrated systems neuroscience, where different multi-channel measures of neural activity are quantitatively related to each other and to computational theory and behavior via the information-rich characterization of distributed representations provided by the RDM. REPRESENTATIONAL SIMILARITY ANALYSIS STEP-BY-STEP In this section we describe the core of RSA step-by-step. We assume that the data to be analyzed consists in a multivariate activity pattern measured for each of a set of conditions in a given brain region, whose representation is to be better understood. The data could be from single-cell or electrode-array recordings, from neuroimaging, or any other modality of brain-activity measurement. We demonstrate the analysis on an fmri experiment, in which human subjects viewed 96 particular object images. The step-bystep description that follows describes the method. The empirical results for our example experiment are described and interpreted subsequently. STEP 1: ESTIMATING THE ACTIVITY PATTERNS The first step of the analysis is the estimation of an activity pattern associated with each experimental condition. In our example, the activity patterns are spatial response patterns from early visual cortex (EVC) and from the fusiform face area (FFA). The analysis proceeds independently for each region. Instead of spatial activity patterns we could use spatiotemporal patterns or simply temporal patterns from a single site as the input to RSA. Similarly, we could filter the measurements in some neuroscientifically meaningful way. For cell recordings, for example, we could use windowed spike counts, multi-unit activity, or local field potentials as the input. In our fmri example, we obtain an activity estimate for each voxel and condition using massively univariate linear modeling (Figure 7). The design matrix used to model each voxel s response is based on the event sequence and a linear model of the hemodynamic response (Boynton et al., 1996). For each region of interest, the resulting condition-related activity patterns form the basis for computation of the representational dissimilarities. STEP 2: MEASURING ACTIVITY-PATTERN DISSIMILARITY In order to compute the RDM (Figure 2), we compare the activity patterns associated with each pair of conditions. A useful measure of activity-pattern dissimilarity that normalizes for both the mean level of activity and the variability of activity is correlation distance, i.e., 1 minus the linear correlation between patterns (cf. Aguirre, 2007; Haxby et al., 2001; Kiani et al., 2007). Alternative measures include the Euclidean distance (cf. Edelman et al., 1998), the Mahalanobis distance (cf. Kriegeskorte et al., 2006) and, in order to relate RSA to conventional activation-based fmri analysis, the absolute value of the regional-average activation difference (Figure 10). The dissimilarity values for all pairs of conditions are assembled in an RDM, which will have a width and height corresponding to the number of conditions and is symmetric about a diagonal of zeros (Figure 2). We can use MDS to visualize the similarity structure of the activity patterns. This is demonstrated in Figure 4, where conditions are represented by colored dots. The distances between the dots approximate the dissimilarities of the activity patterns the conditions are associated with. STEP 3: PREDICTING REPRESENTATIONAL SIMILARITY WITH A RANGE OF MODELS In this section we describe the different types of model that can be evaluated using RSA. Figure 5 shows the internal representations of several example models and Figure 6 shows the dissimilarity matrices characterizing the model representations. Complex computational models In order to evaluate a computational model with RSA, the model needs to simulate some aspect of the information processing occurring in the subject s brain during the experiment. The term model, thus, has a different meaning here than conventionally in statistical data analysis, where it often refers to a statistical model that does not simulate brain information processing (such as the design matrix in Figure 7, which was used to estimate the activity patterns). In our example, we are interested in visual object perception, so the models to be used simulate parts of the visual processing. The models are presented with the same experimental stimuli as our human subjects. Moreover, their internal representations are analyzed in the same way as the measured brain representations of our subjects. We demonstrate RSA with three complex computational models. First, we use a model of V1 consisting in retinotopic maps of simulated simple and complex cells based on banks of Gabor filters for a range of spatial frequencies and orientations at each location (details in the Appendix). We also include a variant of this model, in which we attempted to simulate the local averaging of fmri voxels by pooling local responses of the original V1 model (V1 model, smoothed). Second, as an example of a higher-level representation, we use a model developed in the HMAX framework (Riesenhuber and Poggio, 2002; Serre et al., 2007), which includes C2 units based on natural-image patches as filters and corresponds, approximately, to the level of representation in V4. Third, we use a computational model from computer vision, the RADON transform, whose components in the present implementation Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 6

7 FIGURE 4 Unsupervised arrangement of 96 experimental conditions reflecting pairwise activity-pattern similarity. As in Figure 1, but for 96 instead of 4 conditions, the arrangements reflect the activity-pattern similarity structure. Each panel visualizes the RDM from the corresponding panel in Figure 10. Each condition (corresponding to the presentation of one of 96 object images) is represented by a colored dot, where the color codes for the category (legend at the bottom). In each panel, dots placed close together indicate that the two conditions were associated with similar activity patterns. Dots placed are not meant to resemble neurons in the primate visual system. However, this model could be implemented with biological neurons and has been proposed as a functional account of the representation of visual stimuli in the lateral occipital complex (Wade and Tyler, 2005) based on fmri evidence. Detailed descriptions of the model representations are to be found in the Section Methodological Details. Simple computational models The models described above are meant to simulate brain information processing in some sense. We can additionally use simple image transformations as competing computational models. Although there may be no compelling neuroscientific motivation for such models, they can provide useful benchmarks and help us characterize the information represented in a given brain region. Here we include (1) the digital images themselves in the Lab color space (which more closely reflects human color similarity perception than the RGB color space more commonly used for image storage), (2) the luminance patterns (grayscale versions) of the images, (3) low-pass (i.e., smoothed), and (4) high-pass (i.e., edge-emphasized) versions of the luminance patterns, (5) the Lab joint histograms of the images (representing the set of colors present in each image), and (6) the silhouettes of the objects, in which each figure pixel is 1 and each background pixel 0. These models as well are described in more detail in the Appendix. Conceptual models Model dissimilarity matrices can be obtained not only from explicit computational accounts. A theory may specify that a given brain far apart indicate that the two conditions were associated with dissimilar activity patterns. The panels show results of non-metric multidimensional scaling (minimizing the loss function stress ) for two brain regions (rows) and three activity-pattern dissimilarity measures (columns). Note that a categorical clustering of the face-image response patterns is apparent in the right FFA (bottom row), but not in early visual cortex (top row). (Note that the absolute activation differences could be represented by an arrangement along a straight line, had the dissimilarity matrices not been averaged across subjects.) region represents particular information and abstracts from other information without specifying how the representation is computed. In such conceptual models, the information processing is miraculous (i.e., unspecified) and the activity patterns unknown. However, we can still specify a hypothetical similarity structure to be tested by comparison to the similarity structures found in different brain regions. Here we use two categorical models as examples of this model variety (Figure 6). The first is the animate inanimate model, in which two object images are identical (dissimilarity = 0) if they are either both animate or both inanimate, and different (dissimilarity = 1) if they straddle the category boundary. The second categorical model follows the same logic for the category of faces: two object images are identical (dissimilarity = 0) if they are either both faces or both non-faces, and different (dissimilarity = 1) if exactly one of them is a face. In addition, we use a face-animal-prototype model, which assumes that all faces elicit a prototypical response pattern (implying small dissimilarities between individual face representations) and that the same is true to a lesser degree for the more general class of animal images. Behavior-based similarity structure We could also use behavioral measures to define reference dissimilarity matrices. The dissimilarity values could come from explicit similarity judgments or from reaction times or confusion errors in comparison tasks (Aguirre et al., in preparation; Cutzu and Edelman, 1996, 1998; Edelman et al., 1998; Kiani et al., 2007; Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 7

8 A stimulus image grayscale low-pass grayscale high-pass grayscale silhouette isoluminant V1 simple cell V1 complex cell radon number of pixels RGB intensity B stimulus image grayscale low-pass grayscale high-pass grayscale silhouette isoluminant V1 simple cell V1 complex cell radon FIGURE 5 Model representations of two example images. Two example images (A,B) from the 96-image experiment and their representations in a number of computational models, including standard transformations of image number of pixels RGB intensity processing as well as neuroscientifically motivated models. Note that each such representation defines a unique similarity structure for the 96 stimuli (as encapsulated in the RDMs of Figure 6). Op de Beeck et al. 2001; Shepard et al., 1975). Such behavioral dissimilarity matrices may reflect the representations that determine the behavioral choices, reaction times, or confusion errors. A close match between the RDM of a brain region and the behavioral dissimilar it y mat r ix would suggest that the re g ional representation might play a role in determining the behavior measured. Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 8

9 FIGURE 6 Representational dissimilarity matrices for models and brain regions. Dissimilarity matrices for model representations and regional brain representations (as introduced in Figure 2). The dissimilarity measure is 1 correlation (Pearson correlation across space). Note that each model yields a unique representational similarity structure that can be compared to that of each brain region (bottom five matrices). This comparison is carried out quantitatively in the following figure. The text labels indicate the representation depicted with the color indicating the type: complex computational model (blue), simple computational model (black), conceptual model (green), brain representation (red). Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 9

10 FIGURE 7 Design matrix for condition-rich ungrouped-events fmri design. Both panels illustrate the design matrix used for the 96-image experiment, an example of a condition-rich ungrouped-events design. The top panel shows the hemodynamic predictor time courses for the experimental events occurring in the first couple of minutes of the first run. Note that events occur at 4-s trialonset asynchrony, yielding overlapping but dissociable hemodynamic responses and a reasonable frequency of stimulus presentation. (Each of the 96 conditions occurs exactly once in each run. The condition sequence is independently randomized for each run.) The bottom panel shows the complete design matrix with predictor amplitude color coded (see colorbar on the right). In addition to the 96 predictors for the experimental conditions, the design matrix also includes components modeling slow artefactual trends and residual headmotion artefacts (after rigid-body head-motion correction), and a confound-mean predictor for each run. STEP 4: COMPARING BRAIN AND MODEL DISSIMILARITY MATRICES Once the dissimilarity matrices of the brain representations (Figure 10) and those of theoretical models (Figure 6) have been specified they can be visually and quantitatively compared. One way to quantify the match between two dissimilarity matrices is by means of a correlation coefficient. We use 1-correlation as a measure Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 10

11 of the dissimilarity between RDMs (Figure 8). Because dissimilarity matrices are symmetrical about a diagonal of zeros, the correlation is computed over the values in the upper (or equivalently the lower) triangular region. Note that above we suggested the use of this measure for comparing activity patterns. Here we suggest using it to assess second-order dissimilarity: the dissimilarity of dissimilarity matrices. We could use an alternative distance measure, such as the Euclidean distance, for comparing dissimilarity matrices. As for comparing activity patterns, we again prefer correlation distance, because it is invariant to differences in the mean and variability of the dissimilarities. For the models we use here, we do not wish to assume a linear match between dissimilarity matrices. We therefore use the Spearman rank correlation coefficient to compare them. In the Appendix, we present another argument for the use of rank-correlation distance (instead of the Pearson linear correlation distance or Euclidean distance) for comparing dissimilarity matrices. The argument is based on the observation that, in high-dimensional response spaces, a prominent component of the effect of activity-pattern noise on the dissimilarities can be accounted for by a monotonic transform. Figure 8 shows the deviations (1-Spearman correlation) of the models from each brain region s RDM. Smaller bars indicate better fits. In order to estimate the variability of each model deviation expected if a similar experiment were to be performed with different stimuli (from the same population of stimuli), we computed each model deviation 100 times over for bootstrap resamplings of the condition set (i.e., 96 conditions chosen with replacement from the original set of 96 on each iteration) 4. This method is attractive, (1) because it requires few assumptions, (2) because only the dissimilarity matrices are needed as input, (3) because it is computationally less intensive than modeling the noise at a lower level, and (4) because it generalizes (to the degree possible given the experimental data) from the set of conditions actually used in the experiment to the population of conditions that the actual conditions can be considered a random sample of. This bootstrap procedure would also lend itself to testing whether one model fits the data better than another model, as discussed in the Appendix 5. STEP 5: TESTING RELATEDNESS OF TWO DISSIMILARITY MATRICES BY RANDOMIZATION In order to decide whether two dissimilarity matrices are related, we can perform statistical inference on the RDM correlation. The classical method for testing correlations assumes independent measurements for the two variables. For dissimilarity matrices 4 A complication of this method is that bootstrap resampling of the condition set moves zeros from the diagonal into the off-diagonal parts of the matrix whenever a condition is selected multiple times in the bootstrap resampling. The inclusion of these off-diagonal zeros leads to artefactually small model deviation estimates (because it increases the correlation between the dissimilarity values). In order to avoid underestimating the model deviations in the bootstrap simulation, these artefactual off-diagonal zeros (about 1% of the dissimilarity values here) were excluded before computing the model deviations. 5 Alternatively, we could obtain error bars and statistical tests by estimating the distribution of the model deviation estimates for repetitions of the experiment with the same stimuli and subjects or with the same stimuli and different subjects, or with different stimuli and different subjects. These approaches would provide complementary information to the condition-label bootstrap approach we have described. such independence cannot be assumed, because each similarity is dependent on two response patterns, each of which also codetermines the similarities of all its other pairings in the RDM. We therefore suggest testing the relatedness of dissimilarity matrices by randomizing the condition labels. We choose a random permutation of the conditions, reorder rows and columns of one of the two dissimilarity matrices to be compared according to this permutation, and compute the correlation. Repeating this step many times (e.g., 10,000 times), we obtain a distribution of correlations simulating the null hypothesis that the two dissimilarity matrices are unrelated. If the actual correlation (for consistent labeling between the two dissimilarity matrices) falls within the top α 100% of the simulated null distribution of correlations, we reject the null hypothesis of unrelated dissimilarity matrices with a false-positives rate of α. The p-value for each brain region s relatedness to each model is given beneath the model s bar in Figure 8. They are conservative estimates based on 10,000 random relabelings, so the smallest possible estimate is STEP 6: VISUALIZING THE SIMILARITY STRUCTURE OF REPRESENTATIONAL DISSIMILARITY MATRICES BY MDS MDS provides a general method for arranging entities in a lowdimensional space (e.g., the 2D of a figure on paper), such that their distances reflect their similarities: Similar entities will be placed together, dissimilar entities apart. In Figure 4 we used MDS to visualize the similarity structure of activity patterns in EVC and FFA. Here we suggest using MDS also to visualize the similarity structure of RDM. We first assemble all pairwise comparisons between activitypattern dissimilarity matrices in a dissimilarity matrix of dissimilarity matrices (Figure 9A), using rank-correlation as the dissimilarity measure as suggested above. We then perform MDS on the basis of this second-order dissimilarity matrix. This exploratory visualization technique (Figure 9B) simultaneously relates all RDMs (from models and brain regions) to each other. It thus summarizes the information we would get by inspecting a bar graph of RDM fits (Step 4) not just for EVC and the right FFA (as shown in Figure 8), but for each model and region. The conciseness of the MDS visualization comes at a cost: the distances are distorted (depending on the number of representations included) and there are no error bars or statistical indications. Nevertheless this exploratory visualization technique provides a useful overall view. It can alert us to relationships we had not considered and prompt confirmatory follow-up analysis. EMPIRICAL RESULTS AND THEIR INTERPRETATION THE REPRESENTATIONAL DISSIMILARITY MATRICES OF EVC AND FFA Figure 10 shows that the correlation-distance matrix for EVC and FFA. For the FFA, but not EVC, the matrix reflects the categorical structure of the stimuli. This structure is obvious, because the condition sequence for the dissimilarity matrices were defined by the categorical order. Note, however, that this order affects merely the visual appearance of the matrices. Reordering the conditions does not affect the results of RSA. For the FFA, the correlation-distance matrix reveals a pattern markedly different from that exhibited by the two other measures of activity-pattern dissimilarity. The absolute-activation-difference matrix shows the prominent Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 11

12 FIGURE 8 Matching models to brain regions by comparing representational dissimilarity matrices. The dissimilarity matrices characterizing the representation in early visual cortex (top) and the right FFA (bottom) are compared to dissimilarity matrices obtained from model representations and other brain regions. Each bar indicates the deviation between the RDM of the reference region (early visual cortex or the right FFA) and that of a model or other brain region. The deviation is measured as 1 minus the Spearman correlation between dissimilarity matrices (for motivation see Step 4 and Appendix). Text-label colors indicate the type of representation: complex computational model (blue), simple computational model (black), conceptual model (green), brain representation (red). Error bars indicate the standard error of the deviation estimate. (The standard error is estimated as the standard deviation of 100 deviation estimates obtained from bootstrap resamplings of the conditions set.) The number below each bar indicates the p-value for a test of relatedness between the reference matrix (early visual cortex or the right FFA) and that of the model or other region. (The test is based on 10,000 randomizations of the condition labels.) The black line indicates the noise floor, i.e., the expected deviation between the empirical reference RDM (with noise) and the underlying true RDM (without noise). The red line indicates the expected retest deviation between the empirical dissimilarity matrices that would be obtained for the reference region if the experiment were repeated with different subjects (both matrices affected by noise). Both of these reference lines as well as the dissimilarity signal-to-noise ratios (dissimilarity SNR: below the titles) are estimated from the variability of the dissimilarity estimates across subjects. Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 12

13 A B FIGURE 9 Simultaneously relating all pairs of representations. Figure 8 showed the relationships between two reference regions and all models and other regions. Here we simultaneously visualize the pair-relationships between all models and regions (text labels). Note that the visualization of all pair-relationships comes at a cost: statistical information is omitted here. Textlabel colors indicate the type of representation: complex computational model (blue), simple computational model (black), conceptual model (green), brain representation (red). (A) The correlation matrix (Spearman rank correlation) of RDMs. (B) Multidimensional scaling arrangement (minimizing metric stress) of the representations. Note that MDS was used here to arrange not activity patterns (as in Figures 1 and 4), but dissimilarity matrices. The rubberband graph (gray connections) depicts the inevitable distortions introduced by arranging the models in 2D (see legend of Figure 1 for an explanation). Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 13

14 FIGURE 10 Dissimilarity matrices of activity patterns elicited in early visual cortex and FFA by viewing 96 object images. Dissimilarity matrices (as introduced in Figure 2) are shown for early visual cortex (top row) and right FFA (bottom row) and for three different measures of dissimilarity (columns): 1 correlation (Pearson correlation across space), the Euclidean distance between the two response patterns (in standard error units) and the absolute activation difference (i.e., the absolute value of the difference of the spatial-mean activity level). The absolute activation difference is sensitive only to the overall level of activation and has been included only because regional-average activation is conventionally targeted in fmri analysis. Note that the correlation distance (1 correlation) normalizes for both the overall activation and the variability of activity across space. It is therefore the preferred measure for detecting distributed representations without sensitivity to the global activity level (which could be attributed e.g., to attention). The Euclidean distance combines sensitivity to pattern shape, spatial-mean activity level, and variability across space. Note that as expected using the Euclidean distance yields an RDM resembling both the one obtained with correlation distance and the one obtained with absolute activation difference. The matrices have been separately histogram-equalized (percentile units) for easier comparison. Dissimilarity matrices were averaged across two sessions for each of four subjects. con t r a s t i n a c t iv a t i o n l e ve l b e t we e n f a ce s a n d i n a n i m a te o b j e c t s and less prominently between animate and inanimate objects. The correlation-distance matrix normalizes out the regional-average activation effects and reveals that the activity patterns are highly correlated among faces (human or animal) and to a lesser degree among animals. The Euclidean-distance matrix is sensitive to both the absolute activation difference and the pattern correlation. Unless indicated otherwise, subsequent analyses are based on correlation-distance matrices. THE SIMILARITY STRUCTURE OF ACTIVITY PATTERNS IN EVC AND FFA AS REVEALED BY MDS Figure 4 visualizes the dissimilarity structure as estimated with the three measures by arranging dots that represent the 96 object images in 2D with category-color-coding, such that stimuli eliciting similar response patterns are placed close together and stimuli eliciting dissimilar response patterns are placed far apart. Such arrangements are computed by MDS. We observe some categorical clustering (for faces and, to a lesser degree, for animate objects) in FFA, but not in EVC. This is consistent with our inspection of dissimilarity matrices in Figure 10. MODEL FITS TO EVC AND FFA Figure 6 shows the RDMs of the models. The first thing to note is that each matrix presents a unique pattern that characterizes the model representation. Figure 8 shows the deviation of each model from the empirical RDMs of EVC and FFA. We do not have the space here to fully discuss the neuroscientific implications of this analysis, but we offer some basic observations that demonstrate how RSA can help characterize regional representations: For EVC, note that the best-fitting model is the silhouetteimage model. This is plausible because EVC is known to contain retinotopic representations of the visual input. The fmri patterns in EVC appear to reflect primarily the shape of the retinotopic region stimulated (i.e., the shape of the figure, Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 14

15 since the background is uniformly gray). That the simple silhouette model explains the RDM better here than the V1 model suggests that the orientation information is not as strongly reflected in the RDM. This is consistent with recent results by Kay et al. (2008), who showed that images can be identified on the basis of their fmri responses in EVC, with the major portion of the information provided by the retinotopic representation of edge energy and a smaller portion provided by the representation of edge orientation 6. Early visual orientation information is likely to be attenuated in fmri data because of its fine-scale spatial organization and pooling of columns of all orientation-preferences in each fmri voxel. Among the complex computational models, the V1 model fits the EVC data best, but only the smoothed version, where we simulated local pooling of orientation-specific responses in fmri voxels. Like the good fit of the silhouette model, this is consistent with the limited spatial resolution of our fmri voxels. The RDMs of the fusiform face and parahippocampal place areas in either hemisphere fit the EVC matrix better than the V1 model, but not as well as the silhouette model. One explanation for this is that the conventional V1 model does not capture the full complexity of the representation in EVC. This would be plausible for two reasons: On the one hand, our EVC region contains voxels from the early visual foveal confluence, not just from V1. On the other hand, V1 itself is likely to contain a more complex representation than our Gabor-based model of simple and complex cells. The higher-level HMAX-C2 representation based on natural image patches, plausibly does not capture the similarity structure we find in EVC, nor do the simple image transformations. For the right FFA, the best-fitting dissimilarity structure consists in the empirical dissimilarity of FFA in the opposite hemisphere. This is plausible, given the close functional relationship between the regions. The dissimilarities of the right FFA are best modeled by a conceptual model: the face-animal-prototype model. This suggests that, to a first approximation, different faces elicit a prototypical response pattern implying small dissimilarities between individual face response patterns, consistent with Kriegeskorte et al. (2007), and that the same is true to a lesser degree for the more general class of animal images. 6 Note that Kay et al. (2008) used stimuli of about 20 visual angle (in contrast to the 2.9 stimuli used here) thus driving a more extended retinotopic representation, which may provide more power for detecting the subtler orientation information present in the fmri signals. Note also that the two studies take very different approaches to activity-pattern analysis. Finally, the stimulus set always influences what aspects of a representation we are sensitive to in any neurophysiological experiment. Our stimulus set here may not afford great sensitivity to orientation information in the context of RSA: A given pair of images may be similar in orientation at one retinal location and dissimilar at another, such that the overall representational dissimilarity (across the entire extent of the image) ends up at an intermediate value for all pairs of images. Different results might be expected for grating stimuli, where some stimulus pairs are similar in orientation across the entire extent of the image, and other pairs are dissimilar in orientation everywhere (cf. Kamitani and Tong, 2005) Among the complex computational models, the HMAX-C2 representation based on natural image patches provides the best fit to the right FFA. This may reflect the higher-level nature of the representations in FFA. The right FFA resembles the EVC more closely than the V1 model, the silhouette model, or any other brain region. This could reflect feedback from FFA to EVC. Alternatively, FFA may reflect some of the more complex features of the early visual representation that are not captured by either the silhouette or the V1 model. THE SIMILARITY STRUCTURE OF REPRESENTATIONAL DISSIMILARITY MATRICES AS REVEALED BY MDS Figure 9 simultaneously relates the RDM signatures of all brain regions and models to each other by means of MDS. This representation is devoid of indications of statistical significance and inevitably compromised by geometric distortions (because a higher-dimensional structure is represented in 2D). However, it provides a useful overview of all pairwise relationships (not just the relationships shown in Figure 8 of EVC and the right FFA to the other representations). Although the 2D distances do not precisely reflect the actual dissimilarities between the dissimilarity matrices, almost all observations from Figure 8 are also reflected in the MDS arrangement of Figure 9. However, the MDS arrangement provides us with a lot of additional information. As examples of the additional information, consider these observations: The close interhemispheric observed for the left and right FFA (Figure 8), also holds for the left and right parahippocampal place area. The smoothing applied to the V1 model and the RADON model in order to simulate pooling of responses within fmri voxels does not appear to drastically alter the RDM of either of these models. The five brain regions included (red) all seem to be somewhat related in their representational similarity structure. The fact that no model appears in their midst suggests that there may be a common component to these visual representations that is not captured by any of the models. THE BROAD POTENTIAL OF REPRESENTATIONAL SIMILARITY ANALYSIS RELATING MODELS, BRAIN REGIONS, SUBJECTS, SPECIES, AND BEHAVIOR Systems neuroscience has struggled to quantitatively relate its three major branches of research: behavioral experimentation, brainactivity experimentation, and computational modeling. The RDM can serve as a hub that relates representations from a variety of sources in the three branches (Figure 3). We can use dissimilarity matrices to compare internal representations between two models or two brain regions in the same subject (representational connectivity, see below). In addition, RSA provides a solution to the fine-grained spatial-correspondency problem encountered when relating corresponding brain regions in different subjects of an fmri experiment. Conventionally, different subjects in an fmri experiment are related by transforming the data into a common Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 15

16 spatial frame of reference, such as Talairach space (Talairach and Tournoux, 1988) or cortical-surface space defined by cortex-based alignment (Fischl et al., 1999; Goebel and Singer, 1999; Goebel et al., 2006). However, these available common spaces do not have sufficient precision to relate high-resolution fmri voxels. Establishing spatial correspondency is not merely a technical challenge. It is a fundamental empirical question to what spatial precision intersubject correspondency even exists in different functional areas (Kriegeskorte and Bandettini, 2007). RSA offers an attractive way of abstracting from the spatial layout and even from the linear basis of the representation, allowing us to relate fine-grained activity patterns between subjects. Even different species and modalities of brain-activity data (e.g., single-cell recording and fmri; Kriegeskorte et al., in press) can be meaningfully related with RSA. ADVANCED TYPES OF REPRESENTATIONAL SIMILARITY ANALYSIS Similarity searchlight: Finding brain regions matching a model RSA also allows us to localize a brain region whose intrinsic representation resembles that of a specified model. For this purpose we can move a spherical or cortex-patch searchlight (Kriegeskorte et al., 2006) throughout the measured volume to select, at each location, a local contiguous set of voxels, for which RSA is performed. The results, for each model, form a continuous statistical brain map reflecting how well that model fits in each local neighborhood. Representational connectivity analysis In order to assess to what extent two brain regions in the same subject represent the same information, we can compare the two regions condition- or time-point-based dissimilarity matrices (Kriegeskorte et al., in press). The latter approach can be applied to either the raw data or residuals of the linear modeling of stimulus-related effects. Using the residuals will focus the analysis on the internal representational dynamics of the system including stochastic innovations. In analogy to functional connectivity analysis, we refer to this approach as representational connectivity analysis. It can be combined with the searchlight approach (Kriegeskorte et al., 2006) in order to find a set of regions representationally connected to a given region. Fitting parameters of computational models The computational models we present as examples here are fixed models in that they do not have any parameters fitted on the basis of the data. It will be interesting to extend our approach to the fitting of model parameters on the basis of an empirical RDM. For example, a network model could be trained (supervised learning) to fit a given RDM. In order to avoid circular (i.e., self-fulfilling) inference, a separate set of conditions (e.g., different experimental stimuli) will then be needed to assess the fit of the computational model to the experimental data. Composite modeling of a brain region s representational dissimilarity matrix In our demonstration here, we have treated the models as separate accounts of the data to be evaluated independently. A complementary approach is to model the RDM of a brain region by combining several models. To this end, one could combine units from the internal representations of several models (as we have done for simple and complex V1-model units) and compute the overall representational dissimilarity. One could then fit parameters, including the number of units from each model to include in the representation, so as to best account for an empirical RDM. A simpler approach is to directly model an empirical RDM as a combination of model dissimilarity matrices. If we use Euclidean distance to compare activity patterns and assume that the different models account for orthogonal components of the activity patterns (e.g., separate sets of units), then we can account for the squared empirical Euclidean distance matrix as a linear combination of the squared model Euclidean distance matrices. (Note that this does not require the dissimilarity patterns of the models to be orthogonal; the linear model would use the dissimilarity variance uniquely explained by each model to disambiguate the explanation of shared dissimilarity variance.) A more generally applicable approach would be to explain the empirical RDM as a weighted sum of monotonically transformed model dissimilarity matrices, where a separate monotonic transform is estimated for each model simultaneously with the weights. Weighted representational readout analysis So far we have thought of a region s representation as characterized by a single RDM. Alternatively, we can consider the representation as a high-dimensional structure that is viewed from different perspectives by the regions that read it out. If readout consists in multiple linear weightings of the representational units, then it amounts to a linear projection that can be likened to the transformation of a 3-D structure to a 2-D view of it. In this spirit, we can reverse the logic of the previous paragraph and see to what extent we can read out a particular dissimilarity structure from the representation by weighting the units before computing the RDM. Again, using the squared Euclidean distance yields a simple relationship: Each unit (e.g., a voxel or a neuron) yields a separate RDM. The overall squared Euclidean distance matrix is the sum of the single-unit squared Euclidean distance matrices. Now we can account for each model s dissimilarity pattern as a linear combination of the single-unit dissimilarity matrices. This avenue can be construed as a generalization of linear discriminant analysis from a single contrast to a complex pattern of contrast predictions. It is interesting because of its neuroscientific motivation in terms of readout by other brain regions. As in linear discriminant analysis and classification in general, independent test data will be needed to confirm any relationships suggested by such a fit. CORE CONCEPTS FOR EXPERIMENTAL DESIGN What experimental designs lend themselves to RSA? A distinguishing feature of RSA is its potential to simultaneously exploit the spatial and temporal richness of multi-channel brain-activity data. Although RSA can be applied to a wide range of conventional experimental designs, there may be little conceptual motivation for it in the context of certain experiments, e.g., a low-resolution block-design fmri experiment that targets regional activation and averages across very different processes (e.g., perception of different Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 16

17 stimuli within a given category). The benefits of RSA will be greatest for condition-rich experimental designs targeting activity-pattern information with high-resolution measurement. In this section we describe novel types of experimental design that are feasible with RSA and optimally exploit its potential. Condition-rich design RSA is particularly useful in conjunction with condition-rich designs. One example of such a design is the 96-object-image experiment we presented to demonstrate the approach. We refer to a design as condition-rich if the number of effective experimental conditions (that is brain states to be discerned) is large. Conditionrich designs approach the limit of the temporal complexity of the signal measured in order to amply sample the space of all possible conditions. Within the classical approach of massively univariate activation-based analysis (Friston et al., 1994, 1995; Worsley and Friston, 1995; Worsley et al., 1992), one way of enriching design has been to parameterize the conditions. The result is a larger number of conditions that might not singly yield stable estimates, but the correlation between condition parameters and brain activity combining evidence across conditions can be stably estimated. Such designs also lend themselves to RSA: The model dissimilarity matrices can be computed from the condition parameters. However, RSA is not limited to designs whose conditions sample a predefined parameter space in a regular way. In RSA, the parametric statistical models describing activity variation across time are replaced by computational models exposed to the same experimental conditions. Regular parameterization may help focus the experiment on particular hypotheses, but RSA also accommodates less restricted designs such as the 96-object-image design we use as an example here. Ungrouped-events design In the classical block-design approach to fmri experimentation, an experimental block corresponding to one of the conditions typically includes a variety of brain states (e.g., corresponding to percepts of a variety of stimuli from the same category) that are to be averaged across. While differences between block-average activation can be very sensitively detected with this method, the average results will be ambiguous with respect to single-trial processing (Bedny et al., 2007; Kriegeskorte et al., 2007). Equally importantly, the temporal capacity of the fmri signal to discern a large number of separate brain states is largely wasted. In event-related designs (Buckner, 1998), stimuli can appear in complex temporal sequences allowing for a wider range of experimental tasks. However, the experimental events are usually still grouped in condition sets and the variety of events forming a single condition is averaged across in the analysis (e.g., by modeling each condition by a single predictor). The sequence of experimental events is often designed to maximize estimation efficiency for the condition contrasts of interest. In that case the design itself will imply a grouping of the experimental events. We propose to avoid any predefined grouping of experimental events (ungrouped-events design). Each experimental event (e.g., each stimulus) is treated as a separate condition (Figure 7; Aguirre 2007; Kriegeskorte et al., 2007; Kriegeskorte et al., in press). The 4-image experiment is an example of an ungrouped-events design. The 96-image experiment is an example of an ungrouped events design, which is also condition-rich. One approach is to have events occur in a random sequence implying no grouping. In order to include a reasonable number of events, but still be able to discern the activity patterns they are associated with, we use a design with temporally overlapping but still separable single-trial hemodynamic responses here. Our example employs a design with a trial-onset asynchrony (TOA) of 4 s (Figure 7). The effects of varying the TOA are explored in Figure 11. A more detailed discussion of optimal event sequences for condition-rich designs (including ungrouped-events designs) is to be found in the Appendix (Section Optimal Condition-Rich fmri Design ). For estimation of a given contrast of interest, a condition-rich ungrouped-events design with a random sequence will be less efficient than a block-design or a sequence-optimized rapid eventrelated design. In our view, however, the statistical cost is more than offset by the ability to group the events into arbitrary sets and, more generally, to study the rich space they populate and its relationship to the brain-activity patterns they are associated with. RSA provides an attractive method for exploring this rich empirical information and testing particular hypotheses. Unique-events design and time-continuous experimentation An ungrouped-events design does not group different experimental events into a condition set, but it may contain repetitions of identical experimental events. An extreme type of ungroupedevents design would be a unique-events design, in which no experimental event is ever repeated. RSA can handle uniqueevents designs just like any other design. This is an important property, because unique-events designs take the complexity of the conditions set to the limit of the temporal capacity of the measured signal. In addition, there are neuroscientific domains, where exact repetition of an experimental event is a questionable concept. Strictly speaking each experimental event in any experiment and in fact any experienced event at all permanently changes the brain. In many studies, we may choose a design that minimizes such effects so that we can neglect them in the analysis. For studies of plasticity, however, it may be attractive to track changes to the system along with its activity dynamics. RSA in conjunction with a suitably plastic computational model could address this challenge. We can go one step further and abolish the notion of discrete experimental events in favor of that of time-continuous experimentation (e.g., Hasson et al., 2004). For time-continuous designs, we can treat each acquired volume as a separate condition and directly compute the RDM from the data. For each region of interest, the resulting RDM will then have a width and height corresponding to the number of time points. We refer to such a dissimilarity matrix as a time 2 dissimilarity matrix. For fmri data, the time 2 dissimilarity matrix will reflect the temporal characteristics of the hemodynamic response. Time-continuous RSA is attractive for studies of time-continuous perception of stimuli, including complex natural stimuli such as movies (Hasson et al., 2004), and, more generally, for studies of time-continuous interactions, such as playing computer games Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 17

18 FIGURE 11 Design efficiency as a function of trial-onset asynchrony for a 96-condition fmri design. This figure shows simulation results exploring how statistical efficiency depends on the trial-onset asynchrony (TOA) under linearsystems assumptions for a 96-condition design with one hemodynamicresponse predictor per condition and a random sequence of experimental events (including 25% null events for baseline estimation). We assume that about 50 min of fmri data are to be collected in a single subject. The simulation suggests a simple conclusion: The more closely the trials are spaced in time, the higher the efficiency will be (top panels) for single-conditions (cyan) and pairwise condition contrasts (red). Doubling the number of trials packed into the same 50- min period, then, would improve efficiency about as much as performing the whole experiment twice: decreasing the standard errors of the estimates roughly by a factor of sqrt(2). In other words, the standard errors are proportional to sqrt(toa). (Why does not the greater response overlap decrease efficiency? For an intuitive understanding, consider that although the greater response overlap for shorter TOAs correlates predictors, the greater number of event repetitions decorrelates them.) Importantly, however, the straightforward relationship suggested by the simulation rests on the assumption of a linear neuronal and hemodynamic response system. In reality, the effects of closely spaced events may interact at the neuronal level and the hemodynamic responses may also not behave linearly (e.g., three 16-ms stimuli at a TOA of 32- ms are unlikely to elicit a hemodynamic response that is three times higher than that to a single such stimulus). The choice of TOA therefore requires an informed guess regarding the short-toa nonlinearity for the particular experimental events used. For the 96-image experiment, we chose a TOA of 4 s. Details on the simulation and an intuitive explanation for the result are given in the Appendix (Section Optimal Condition-Rich fmri Design ), along with further discussion of design choices including the TOA. Frontiers in Systems Neuroscience November 2008 Volume 2 Article 4 18

Matching Categorical Object Representations in Inferior Temporal Cortex of Man and Monkey

Matching Categorical Object Representations in Inferior Temporal Cortex of Man and Monkey (2008), doi:10.1016/j.neuron.2008.10.043 Article Matching Categorical Object Representations in Inferior Temporal Cortex of Man and Monkey Nikolaus Kriegeskorte, 1, * Marieke Mur, 1,2 Douglas A. Ruff,

More information

Matching Categorical Object Representations in Inferior Temporal Cortex of Man and Monkey

Matching Categorical Object Representations in Inferior Temporal Cortex of Man and Monkey Article Matching Categorical Object Representations in Inferior Temporal Cortex of Man and Monkey Nikolaus Kriegeskorte, 1, * Marieke Mur, 1,2 Douglas A. Ruff, 1 Roozbeh Kiani, 3 Jerzy Bodurka, 1,4 Hossein

More information

Representational similarity analysis

Representational similarity analysis School of Psychology Representational similarity analysis Dr Ian Charest Representational similarity analysis representational dissimilarity matrices (RDMs) stimulus (e.g. images, sounds, other experimental

More information

Pattern-information analysis: from stimulus decoding to computational-model testing

Pattern-information analysis: from stimulus decoding to computational-model testing Pattern-information analysis: from stimulus decoding to computational-model testing Nikolaus Kriegeskorte Medical Research Council, Cognition and Brain Sciences Unit, 15 Chaucer Road, Cambridge, CB2 7EF,

More information

SUPPLEMENTARY INFORMATION. Table 1 Patient characteristics Preoperative. language testing

SUPPLEMENTARY INFORMATION. Table 1 Patient characteristics Preoperative. language testing Categorical Speech Representation in the Human Superior Temporal Gyrus Edward F. Chang, Jochem W. Rieger, Keith D. Johnson, Mitchel S. Berger, Nicholas M. Barbaro, Robert T. Knight SUPPLEMENTARY INFORMATION

More information

Prediction of Successful Memory Encoding from fmri Data

Prediction of Successful Memory Encoding from fmri Data Prediction of Successful Memory Encoding from fmri Data S.K. Balci 1, M.R. Sabuncu 1, J. Yoo 2, S.S. Ghosh 3, S. Whitfield-Gabrieli 2, J.D.E. Gabrieli 2 and P. Golland 1 1 CSAIL, MIT, Cambridge, MA, USA

More information

Supporting Information

Supporting Information 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 Supporting Information Variances and biases of absolute distributions were larger in the 2-line

More information

Sum of Neurally Distinct Stimulus- and Task-Related Components.

Sum of Neurally Distinct Stimulus- and Task-Related Components. SUPPLEMENTARY MATERIAL for Cardoso et al. 22 The Neuroimaging Signal is a Linear Sum of Neurally Distinct Stimulus- and Task-Related Components. : Appendix: Homogeneous Linear ( Null ) and Modified Linear

More information

Supplemental Material

Supplemental Material 1 Supplemental Material Golomb, J.D, and Kanwisher, N. (2012). Higher-level visual cortex represents retinotopic, not spatiotopic, object location. Cerebral Cortex. Contents: - Supplemental Figures S1-S3

More information

Deep Neural Networks Rival the Representation of Primate IT Cortex for Core Visual Object Recognition

Deep Neural Networks Rival the Representation of Primate IT Cortex for Core Visual Object Recognition Deep Neural Networks Rival the Representation of Primate IT Cortex for Core Visual Object Recognition Charles F. Cadieu, Ha Hong, Daniel L. K. Yamins, Nicolas Pinto, Diego Ardila, Ethan A. Solomon, Najib

More information

Local Image Structures and Optic Flow Estimation

Local Image Structures and Optic Flow Estimation Local Image Structures and Optic Flow Estimation Sinan KALKAN 1, Dirk Calow 2, Florentin Wörgötter 1, Markus Lappe 2 and Norbert Krüger 3 1 Computational Neuroscience, Uni. of Stirling, Scotland; {sinan,worgott}@cn.stir.ac.uk

More information

Word Length Processing via Region-to-Region Connectivity

Word Length Processing via Region-to-Region Connectivity Word Length Processing via Region-to-Region Connectivity Mariya Toneva Machine Learning Department, Neural Computation Carnegie Mellon University Data Analysis Project DAP committee: Tom Mitchell, Robert

More information

Chapter 1. Introduction

Chapter 1. Introduction Chapter 1 Introduction 1.1 Motivation and Goals The increasing availability and decreasing cost of high-throughput (HT) technologies coupled with the availability of computational tools and data form a

More information

Morton-Style Factorial Coding of Color in Primary Visual Cortex

Morton-Style Factorial Coding of Color in Primary Visual Cortex Morton-Style Factorial Coding of Color in Primary Visual Cortex Javier R. Movellan Institute for Neural Computation University of California San Diego La Jolla, CA 92093-0515 movellan@inc.ucsd.edu Thomas

More information

A contrast paradox in stereopsis, motion detection and vernier acuity

A contrast paradox in stereopsis, motion detection and vernier acuity A contrast paradox in stereopsis, motion detection and vernier acuity S. B. Stevenson *, L. K. Cormack Vision Research 40, 2881-2884. (2000) * University of Houston College of Optometry, Houston TX 77204

More information

Virtual Brain Reading: A Connectionist Approach to Understanding fmri

Virtual Brain Reading: A Connectionist Approach to Understanding fmri Virtual Brain Reading: A Connectionist Approach to Understanding fmri Rosemary A. Cowell (r.a.cowell@kent.ac.uk) Computing Laboratory, University of Kent, Canterbury, CT2 7NF, UK David E. Huber (dhuber@psy.ucsd.edu)

More information

Analysis of in-vivo extracellular recordings. Ryan Morrill Bootcamp 9/10/2014

Analysis of in-vivo extracellular recordings. Ryan Morrill Bootcamp 9/10/2014 Analysis of in-vivo extracellular recordings Ryan Morrill Bootcamp 9/10/2014 Goals for the lecture Be able to: Conceptually understand some of the analysis and jargon encountered in a typical (sensory)

More information

Information Processing During Transient Responses in the Crayfish Visual System

Information Processing During Transient Responses in the Crayfish Visual System Information Processing During Transient Responses in the Crayfish Visual System Christopher J. Rozell, Don. H. Johnson and Raymon M. Glantz Department of Electrical & Computer Engineering Department of

More information

Introduction to Computational Neuroscience

Introduction to Computational Neuroscience Introduction to Computational Neuroscience Lecture 5: Data analysis II Lesson Title 1 Introduction 2 Structure and Function of the NS 3 Windows to the Brain 4 Data analysis 5 Data analysis II 6 Single

More information

Hebbian Plasticity for Improving Perceptual Decisions

Hebbian Plasticity for Improving Perceptual Decisions Hebbian Plasticity for Improving Perceptual Decisions Tsung-Ren Huang Department of Psychology, National Taiwan University trhuang@ntu.edu.tw Abstract Shibata et al. reported that humans could learn to

More information

Visual Categorization: How the Monkey Brain Does It

Visual Categorization: How the Monkey Brain Does It Visual Categorization: How the Monkey Brain Does It Ulf Knoblich 1, Maximilian Riesenhuber 1, David J. Freedman 2, Earl K. Miller 2, and Tomaso Poggio 1 1 Center for Biological and Computational Learning,

More information

Framework for Comparative Research on Relational Information Displays

Framework for Comparative Research on Relational Information Displays Framework for Comparative Research on Relational Information Displays Sung Park and Richard Catrambone 2 School of Psychology & Graphics, Visualization, and Usability Center (GVU) Georgia Institute of

More information

Unit 1 Exploring and Understanding Data

Unit 1 Exploring and Understanding Data Unit 1 Exploring and Understanding Data Area Principle Bar Chart Boxplot Conditional Distribution Dotplot Empirical Rule Five Number Summary Frequency Distribution Frequency Polygon Histogram Interquartile

More information

Retinotopy & Phase Mapping

Retinotopy & Phase Mapping Retinotopy & Phase Mapping Fani Deligianni B. A. Wandell, et al. Visual Field Maps in Human Cortex, Neuron, 56(2):366-383, 2007 Retinotopy Visual Cortex organised in visual field maps: Nearby neurons have

More information

Experimental Design I

Experimental Design I Experimental Design I Topics What questions can we ask (intelligently) in fmri Basic assumptions in isolating cognitive processes and comparing conditions General design strategies A few really cool experiments

More information

Chapter 1: Explaining Behavior

Chapter 1: Explaining Behavior Chapter 1: Explaining Behavior GOAL OF SCIENCE is to generate explanations for various puzzling natural phenomenon. - Generate general laws of behavior (psychology) RESEARCH: principle method for acquiring

More information

MULTIPLE LINEAR REGRESSION 24.1 INTRODUCTION AND OBJECTIVES OBJECTIVES

MULTIPLE LINEAR REGRESSION 24.1 INTRODUCTION AND OBJECTIVES OBJECTIVES 24 MULTIPLE LINEAR REGRESSION 24.1 INTRODUCTION AND OBJECTIVES In the previous chapter, simple linear regression was used when you have one independent variable and one dependent variable. This chapter

More information

Disparity- and velocity- based signals for 3D motion perception in human MT+ Bas Rokers, Lawrence K. Cormack, and Alexander C. Huk

Disparity- and velocity- based signals for 3D motion perception in human MT+ Bas Rokers, Lawrence K. Cormack, and Alexander C. Huk Disparity- and velocity- based signals for 3D motion perception in human MT+ Bas Rokers, Lawrence K. Cormack, and Alexander C. Huk Supplementary Materials fmri response (!% BOLD) ).5 CD versus STS 1 wedge

More information

Advanced Data Modelling & Inference

Advanced Data Modelling & Inference Edinburgh 2015: biennial SPM course Advanced Data Modelling & Inference Cyril Pernet Centre for Clinical Brain Sciences (CCBS) Neuroimaging Sciences Modelling? Y = XB + e here is my model B might be biased

More information

The Integration of Features in Visual Awareness : The Binding Problem. By Andrew Laguna, S.J.

The Integration of Features in Visual Awareness : The Binding Problem. By Andrew Laguna, S.J. The Integration of Features in Visual Awareness : The Binding Problem By Andrew Laguna, S.J. Outline I. Introduction II. The Visual System III. What is the Binding Problem? IV. Possible Theoretical Solutions

More information

Learning to classify integral-dimension stimuli

Learning to classify integral-dimension stimuli Psychonomic Bulletin & Review 1996, 3 (2), 222 226 Learning to classify integral-dimension stimuli ROBERT M. NOSOFSKY Indiana University, Bloomington, Indiana and THOMAS J. PALMERI Vanderbilt University,

More information

CS/NEUR125 Brains, Minds, and Machines. Due: Friday, April 14

CS/NEUR125 Brains, Minds, and Machines. Due: Friday, April 14 CS/NEUR125 Brains, Minds, and Machines Assignment 5: Neural mechanisms of object-based attention Due: Friday, April 14 This Assignment is a guided reading of the 2014 paper, Neural Mechanisms of Object-Based

More information

Neuroimaging. BIE601 Advanced Biological Engineering Dr. Boonserm Kaewkamnerdpong Biological Engineering Program, KMUTT. Human Brain Mapping

Neuroimaging. BIE601 Advanced Biological Engineering Dr. Boonserm Kaewkamnerdpong Biological Engineering Program, KMUTT. Human Brain Mapping 11/8/2013 Neuroimaging N i i BIE601 Advanced Biological Engineering Dr. Boonserm Kaewkamnerdpong Biological Engineering Program, KMUTT 2 Human Brain Mapping H Human m n brain br in m mapping ppin can nb

More information

Caveats for the Spatial Arrangement Method: Supplementary Materials

Caveats for the Spatial Arrangement Method: Supplementary Materials Caveats for the Spatial Arrangement Method: Supplementary Materials Steven Verheyen, Wouter Voorspoels, Wolf Vanpaemel, & Gert Storms University of Leuven, Leuven, Belgium 2 Data Our comment is based on

More information

Neural Correlates of Human Cognitive Function:

Neural Correlates of Human Cognitive Function: Neural Correlates of Human Cognitive Function: A Comparison of Electrophysiological and Other Neuroimaging Approaches Leun J. Otten Institute of Cognitive Neuroscience & Department of Psychology University

More information

Goodness of Pattern and Pattern Uncertainty 1

Goodness of Pattern and Pattern Uncertainty 1 J'OURNAL OF VERBAL LEARNING AND VERBAL BEHAVIOR 2, 446-452 (1963) Goodness of Pattern and Pattern Uncertainty 1 A visual configuration, or pattern, has qualities over and above those which can be specified

More information

Rajeev Raizada: Statement of research interests

Rajeev Raizada: Statement of research interests Rajeev Raizada: Statement of research interests Overall goal: explore how the structure of neural representations gives rise to behavioural abilities and disabilities There tends to be a split in the field

More information

Reveal Relationships in Categorical Data

Reveal Relationships in Categorical Data SPSS Categories 15.0 Specifications Reveal Relationships in Categorical Data Unleash the full potential of your data through perceptual mapping, optimal scaling, preference scaling, and dimension reduction

More information

Inferential Brain Mapping

Inferential Brain Mapping Inferential Brain Mapping Cyril Pernet Centre for Clinical Brain Sciences (CCBS) Neuroimaging Sciences Edinburgh Biennial SPM course 2019 @CyrilRPernet Overview Statistical inference is the process of

More information

Just One View: Invariances in Inferotemporal Cell Tuning

Just One View: Invariances in Inferotemporal Cell Tuning Just One View: Invariances in Inferotemporal Cell Tuning Maximilian Riesenhuber Tomaso Poggio Center for Biological and Computational Learning and Department of Brain and Cognitive Sciences Massachusetts

More information

Convergence Principles: Information in the Answer

Convergence Principles: Information in the Answer Convergence Principles: Information in the Answer Sets of Some Multiple-Choice Intelligence Tests A. P. White and J. E. Zammarelli University of Durham It is hypothesized that some common multiplechoice

More information

Differences of Face and Object Recognition in Utilizing Early Visual Information

Differences of Face and Object Recognition in Utilizing Early Visual Information Differences of Face and Object Recognition in Utilizing Early Visual Information Peter Kalocsai and Irving Biederman Department of Psychology and Computer Science University of Southern California Los

More information

Introduction to MVPA. Alexandra Woolgar 16/03/10

Introduction to MVPA. Alexandra Woolgar 16/03/10 Introduction to MVPA Alexandra Woolgar 16/03/10 MVP...what? Multi-Voxel Pattern Analysis (MultiVariate Pattern Analysis) * Overview Why bother? Different approaches Basics of designing experiments and

More information

fmri: What Does It Measure?

fmri: What Does It Measure? fmri: What Does It Measure? Psychology 355: Cognitive Psychology Instructor: John Miyamoto 04/02/2018: Lecture 02-1 Note: This Powerpoint presentation may contain macros that I wrote to help me create

More information

Theta sequences are essential for internally generated hippocampal firing fields.

Theta sequences are essential for internally generated hippocampal firing fields. Theta sequences are essential for internally generated hippocampal firing fields. Yingxue Wang, Sandro Romani, Brian Lustig, Anthony Leonardo, Eva Pastalkova Supplementary Materials Supplementary Modeling

More information

What Evidence Supports Special Processing for Faces? A Cautionary Tale for fmri Interpretation

What Evidence Supports Special Processing for Faces? A Cautionary Tale for fmri Interpretation What Evidence Supports Special Processing for Faces? A Cautionary Tale for fmri Interpretation Rosemary A. Cowell 1 and Garrison W. Cottrell 2 Abstract We trained a neurocomputational model on six categories

More information

Decoding the Brain: Neural Representation and the Limits of Multivariate Pattern Analysis in Cognitive Neuroscience Abstract 1.

Decoding the Brain: Neural Representation and the Limits of Multivariate Pattern Analysis in Cognitive Neuroscience Abstract 1. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 Decoding the Brain: Neural Representation and the Limits of Multivariate Pattern Analysis in

More information

Analysis of Environmental Data Conceptual Foundations: En viro n m e n tal Data

Analysis of Environmental Data Conceptual Foundations: En viro n m e n tal Data Analysis of Environmental Data Conceptual Foundations: En viro n m e n tal Data 1. Purpose of data collection...................................................... 2 2. Samples and populations.......................................................

More information

TESTING A NEW THEORY OF PSYCHOPHYSICAL SCALING: TEMPORAL LOUDNESS INTEGRATION

TESTING A NEW THEORY OF PSYCHOPHYSICAL SCALING: TEMPORAL LOUDNESS INTEGRATION TESTING A NEW THEORY OF PSYCHOPHYSICAL SCALING: TEMPORAL LOUDNESS INTEGRATION Karin Zimmer, R. Duncan Luce and Wolfgang Ellermeier Institut für Kognitionsforschung der Universität Oldenburg, Germany Institute

More information

Incorporation of Imaging-Based Functional Assessment Procedures into the DICOM Standard Draft version 0.1 7/27/2011

Incorporation of Imaging-Based Functional Assessment Procedures into the DICOM Standard Draft version 0.1 7/27/2011 Incorporation of Imaging-Based Functional Assessment Procedures into the DICOM Standard Draft version 0.1 7/27/2011 I. Purpose Drawing from the profile development of the QIBA-fMRI Technical Committee,

More information

Classification and Statistical Analysis of Auditory FMRI Data Using Linear Discriminative Analysis and Quadratic Discriminative Analysis

Classification and Statistical Analysis of Auditory FMRI Data Using Linear Discriminative Analysis and Quadratic Discriminative Analysis International Journal of Innovative Research in Computer Science & Technology (IJIRCST) ISSN: 2347-5552, Volume-2, Issue-6, November-2014 Classification and Statistical Analysis of Auditory FMRI Data Using

More information

Introduction to Computational Neuroscience

Introduction to Computational Neuroscience Introduction to Computational Neuroscience Lecture 11: Attention & Decision making Lesson Title 1 Introduction 2 Structure and Function of the NS 3 Windows to the Brain 4 Data analysis 5 Data analysis

More information

Prof. Greg Francis 7/31/15

Prof. Greg Francis 7/31/15 s PSY 200 Greg Francis Lecture 06 How do you recognize your grandmother? Action potential With enough excitatory input, a cell produces an action potential that sends a signal down its axon to other cells

More information

SUPPLEMENTAL MATERIAL

SUPPLEMENTAL MATERIAL 1 SUPPLEMENTAL MATERIAL Response time and signal detection time distributions SM Fig. 1. Correct response time (thick solid green curve) and error response time densities (dashed red curve), averaged across

More information

Object recognition and hierarchical computation

Object recognition and hierarchical computation Object recognition and hierarchical computation Challenges in object recognition. Fukushima s Neocognitron View-based representations of objects Poggio s HMAX Forward and Feedback in visual hierarchy Hierarchical

More information

Carnegie Mellon University Annual Progress Report: 2011 Formula Grant

Carnegie Mellon University Annual Progress Report: 2011 Formula Grant Carnegie Mellon University Annual Progress Report: 2011 Formula Grant Reporting Period January 1, 2012 June 30, 2012 Formula Grant Overview The Carnegie Mellon University received $943,032 in formula funds

More information

Plasticity of Cerebral Cortex in Development

Plasticity of Cerebral Cortex in Development Plasticity of Cerebral Cortex in Development Jessica R. Newton and Mriganka Sur Department of Brain & Cognitive Sciences Picower Center for Learning & Memory Massachusetts Institute of Technology Cambridge,

More information

ANALYSIS AND DETECTION OF BRAIN TUMOUR USING IMAGE PROCESSING TECHNIQUES

ANALYSIS AND DETECTION OF BRAIN TUMOUR USING IMAGE PROCESSING TECHNIQUES ANALYSIS AND DETECTION OF BRAIN TUMOUR USING IMAGE PROCESSING TECHNIQUES P.V.Rohini 1, Dr.M.Pushparani 2 1 M.Phil Scholar, Department of Computer Science, Mother Teresa women s university, (India) 2 Professor

More information

Supplementary materials for: Executive control processes underlying multi- item working memory

Supplementary materials for: Executive control processes underlying multi- item working memory Supplementary materials for: Executive control processes underlying multi- item working memory Antonio H. Lara & Jonathan D. Wallis Supplementary Figure 1 Supplementary Figure 1. Behavioral measures of

More information

From feedforward vision to natural vision: The impact of free viewing and clutter on monkey inferior temporal object representations

From feedforward vision to natural vision: The impact of free viewing and clutter on monkey inferior temporal object representations From feedforward vision to natural vision: The impact of free viewing and clutter on monkey inferior temporal object representations James DiCarlo The McGovern Institute for Brain Research Department of

More information

9 research designs likely for PSYC 2100

9 research designs likely for PSYC 2100 9 research designs likely for PSYC 2100 1) 1 factor, 2 levels, 1 group (one group gets both treatment levels) related samples t-test (compare means of 2 levels only) 2) 1 factor, 2 levels, 2 groups (one

More information

Hierarchical Bayesian Modeling of Individual Differences in Texture Discrimination

Hierarchical Bayesian Modeling of Individual Differences in Texture Discrimination Hierarchical Bayesian Modeling of Individual Differences in Texture Discrimination Timothy N. Rubin (trubin@uci.edu) Michael D. Lee (mdlee@uci.edu) Charles F. Chubb (cchubb@uci.edu) Department of Cognitive

More information

Intrinsic Signal Optical Imaging

Intrinsic Signal Optical Imaging Intrinsic Signal Optical Imaging Introduction Intrinsic signal optical imaging (ISOI) is a technique used to map dynamics in single cells, brain slices and even and most importantly entire mammalian brains.

More information

CHAPTER VI RESEARCH METHODOLOGY

CHAPTER VI RESEARCH METHODOLOGY CHAPTER VI RESEARCH METHODOLOGY 6.1 Research Design Research is an organized, systematic, data based, critical, objective, scientific inquiry or investigation into a specific problem, undertaken with the

More information

Chapter 7: Descriptive Statistics

Chapter 7: Descriptive Statistics Chapter Overview Chapter 7 provides an introduction to basic strategies for describing groups statistically. Statistical concepts around normal distributions are discussed. The statistical procedures of

More information

Interpreting fmri Decoding Weights: Additional Considerations

Interpreting fmri Decoding Weights: Additional Considerations Interpreting fmri Decoding Weights: Additional Considerations P.K. Douglas Modeling and Simulation Department,UCF Department of Psychiatry and Biobehavioral Medicine, UCLA Los Angeles, CA 90024 pdouglas@ist.ucf.edu

More information

Lecture overview. What hypothesis to test in the fly? Quantitative data collection Visual physiology conventions ( Methods )

Lecture overview. What hypothesis to test in the fly? Quantitative data collection Visual physiology conventions ( Methods ) Lecture overview What hypothesis to test in the fly? Quantitative data collection Visual physiology conventions ( Methods ) 1 Lecture overview What hypothesis to test in the fly? Quantitative data collection

More information

Functional Elements and Networks in fmri

Functional Elements and Networks in fmri Functional Elements and Networks in fmri Jarkko Ylipaavalniemi 1, Eerika Savia 1,2, Ricardo Vigário 1 and Samuel Kaski 1,2 1- Helsinki University of Technology - Adaptive Informatics Research Centre 2-

More information

Beyond Blind Averaging: Analyzing Event-Related Brain Dynamics. Scott Makeig. sccn.ucsd.edu

Beyond Blind Averaging: Analyzing Event-Related Brain Dynamics. Scott Makeig. sccn.ucsd.edu Beyond Blind Averaging: Analyzing Event-Related Brain Dynamics Scott Makeig Institute for Neural Computation University of California San Diego La Jolla CA sccn.ucsd.edu Talk given at the EEG/MEG course

More information

Chapter 5. Summary and Conclusions! 131

Chapter 5. Summary and Conclusions! 131 ! Chapter 5 Summary and Conclusions! 131 Chapter 5!!!! Summary of the main findings The present thesis investigated the sensory representation of natural sounds in the human auditory cortex. Specifically,

More information

Lecturer: Rob van der Willigen 11/9/08

Lecturer: Rob van der Willigen 11/9/08 Auditory Perception - Detection versus Discrimination - Localization versus Discrimination - - Electrophysiological Measurements Psychophysical Measurements Three Approaches to Researching Audition physiology

More information

Principals of Object Perception

Principals of Object Perception Principals of Object Perception Elizabeth S. Spelke COGNITIVE SCIENCE 14, 29-56 (1990) Cornell University Summary Infants perceive object by analyzing tree-dimensional surface arrangements and motions.

More information

Neural Correlation of Global-first Topological Perception: Anterior Temporal Lobe

Neural Correlation of Global-first Topological Perception: Anterior Temporal Lobe Brain Imaging and Behavior (28) 2:39 317 DOI 1.17/s11682-8-939-y Neural Correlation of Global-first Topological Perception: Anterior Temporal Lobe Tiangang Zhou & Jun Zhang & Lin Chen Received: 8 April

More information

TMS Disruption of Time Encoding in Human Primary Visual Cortex Molly Bryan Beauchamp Lab

TMS Disruption of Time Encoding in Human Primary Visual Cortex Molly Bryan Beauchamp Lab TMS Disruption of Time Encoding in Human Primary Visual Cortex Molly Bryan Beauchamp Lab This report details my summer research project for the REU Theoretical and Computational Neuroscience program as

More information

Lecturer: Rob van der Willigen 11/9/08

Lecturer: Rob van der Willigen 11/9/08 Auditory Perception - Detection versus Discrimination - Localization versus Discrimination - Electrophysiological Measurements - Psychophysical Measurements 1 Three Approaches to Researching Audition physiology

More information

Multivariate Patterns in Object-Selective Cortex Dissociate Perceptual and Physical Shape Similarity

Multivariate Patterns in Object-Selective Cortex Dissociate Perceptual and Physical Shape Similarity Multivariate Patterns in Object-Selective Cortex Dissociate Perceptual and Physical Shape Similarity Johannes Haushofer 1,2*, Margaret S. Livingstone 1, Nancy Kanwisher 2 PLoS BIOLOGY 1 Department of Neurobiology,

More information

Still important ideas

Still important ideas Readings: OpenStax - Chapters 1 13 & Appendix D & E (online) Plous Chapters 17 & 18 - Chapter 17: Social Influences - Chapter 18: Group Judgments and Decisions Still important ideas Contrast the measurement

More information

Dynamics of Color Category Formation and Boundaries

Dynamics of Color Category Formation and Boundaries Dynamics of Color Category Formation and Boundaries Stephanie Huette* Department of Psychology, University of Memphis, Memphis, TN Definition Dynamics of color boundaries is broadly the area that characterizes

More information

Supplementary Information

Supplementary Information Supplementary Information The neural correlates of subjective value during intertemporal choice Joseph W. Kable and Paul W. Glimcher a 10 0 b 10 0 10 1 10 1 Discount rate k 10 2 Discount rate k 10 2 10

More information

Gathering and Repetition of the Elements in an Image Affect the Perception of Order and Disorder

Gathering and Repetition of the Elements in an Image Affect the Perception of Order and Disorder International Journal of Affective Engineering Vol.13 No.3 pp.167-173 (2014) ORIGINAL ARTICLE Gathering and Repetition of the Elements in an Image Affect the Perception of Order and Disorder Yusuke MATSUDA

More information

M Cells. Why parallel pathways? P Cells. Where from the retina? Cortical visual processing. Announcements. Main visual pathway from retina to V1

M Cells. Why parallel pathways? P Cells. Where from the retina? Cortical visual processing. Announcements. Main visual pathway from retina to V1 Announcements exam 1 this Thursday! review session: Wednesday, 5:00-6:30pm, Meliora 203 Bryce s office hours: Wednesday, 3:30-5:30pm, Gleason https://www.youtube.com/watch?v=zdw7pvgz0um M Cells M cells

More information

arxiv: v1 [q-bio.nc] 12 Jun 2014

arxiv: v1 [q-bio.nc] 12 Jun 2014 1 arxiv:1406.3284v1 [q-bio.nc] 12 Jun 2014 Deep Neural Networks Rival the Representation of Primate IT Cortex for Core Visual Object Recognition Charles F. Cadieu 1,, Ha Hong 1,2, Daniel L. K. Yamins 1,

More information

EARLY STAGE DIAGNOSIS OF LUNG CANCER USING CT-SCAN IMAGES BASED ON CELLULAR LEARNING AUTOMATE

EARLY STAGE DIAGNOSIS OF LUNG CANCER USING CT-SCAN IMAGES BASED ON CELLULAR LEARNING AUTOMATE EARLY STAGE DIAGNOSIS OF LUNG CANCER USING CT-SCAN IMAGES BASED ON CELLULAR LEARNING AUTOMATE SAKTHI NEELA.P.K Department of M.E (Medical electronics) Sengunthar College of engineering Namakkal, Tamilnadu,

More information

Analysis and Interpretation of Data Part 1

Analysis and Interpretation of Data Part 1 Analysis and Interpretation of Data Part 1 DATA ANALYSIS: PRELIMINARY STEPS 1. Editing Field Edit Completeness Legibility Comprehensibility Consistency Uniformity Central Office Edit 2. Coding Specifying

More information

Psychology of Perception Psychology 4165, Spring 2015 Laboratory 1 Noisy Representations: The Oblique Effect

Psychology of Perception Psychology 4165, Spring 2015 Laboratory 1 Noisy Representations: The Oblique Effect Psychology 4165, Laboratory 1 Noisy Representations: The Oblique Effect Probability of Clockwise Response 0.0 0.4 0.8 Orientation Discrimination at 0.0 Deg 0 deg -10-5 0 5 10 Orientation of Test Stimulus

More information

Lateral Geniculate Nucleus (LGN)

Lateral Geniculate Nucleus (LGN) Lateral Geniculate Nucleus (LGN) What happens beyond the retina? What happens in Lateral Geniculate Nucleus (LGN)- 90% flow Visual cortex Information Flow Superior colliculus 10% flow Slide 2 Information

More information

Frank Tong. Department of Psychology Green Hall Princeton University Princeton, NJ 08544

Frank Tong. Department of Psychology Green Hall Princeton University Princeton, NJ 08544 Frank Tong Department of Psychology Green Hall Princeton University Princeton, NJ 08544 Office: Room 3-N-2B Telephone: 609-258-2652 Fax: 609-258-1113 Email: ftong@princeton.edu Graduate School Applicants

More information

Consciousness The final frontier!

Consciousness The final frontier! Consciousness The final frontier! How to Define it??? awareness perception - automatic and controlled memory - implicit and explicit ability to tell us about experiencing it attention. And the bottleneck

More information

EDGE DETECTION. Edge Detectors. ICS 280: Visual Perception

EDGE DETECTION. Edge Detectors. ICS 280: Visual Perception EDGE DETECTION Edge Detectors Slide 2 Convolution & Feature Detection Slide 3 Finds the slope First derivative Direction dependent Need many edge detectors for all orientation Second order derivatives

More information

Early Stages of Vision Might Explain Data to Information Transformation

Early Stages of Vision Might Explain Data to Information Transformation Early Stages of Vision Might Explain Data to Information Transformation Baran Çürüklü Department of Computer Science and Engineering Mälardalen University Västerås S-721 23, Sweden Abstract. In this paper

More information

Supplemental Information

Supplemental Information Current Biology, Volume 22 Supplemental Information The Neural Correlates of Crowding-Induced Changes in Appearance Elaine J. Anderson, Steven C. Dakin, D. Samuel Schwarzkopf, Geraint Rees, and John Greenwood

More information

Statistical Methods and Reasoning for the Clinical Sciences

Statistical Methods and Reasoning for the Clinical Sciences Statistical Methods and Reasoning for the Clinical Sciences Evidence-Based Practice Eiki B. Satake, PhD Contents Preface Introduction to Evidence-Based Statistics: Philosophical Foundation and Preliminaries

More information

Attention modulates spatial priority maps in the human occipital, parietal and frontal cortices

Attention modulates spatial priority maps in the human occipital, parietal and frontal cortices Attention modulates spatial priority maps in the human occipital, parietal and frontal cortices Thomas C Sprague 1 & John T Serences 1,2 Computational theories propose that attention modulates the topographical

More information

Evaluation of a Shape-Based Model of Human Face Discrimination Using fmri and Behavioral Techniques

Evaluation of a Shape-Based Model of Human Face Discrimination Using fmri and Behavioral Techniques Neuron 50, 159 172, April 6, 2006 ª2006 Elsevier Inc. DOI 10.1016/j.neuron.2006.03.012 Evaluation of a Shape-Based Model of Human Face Discrimination Using fmri and Behavioral Techniques Xiong Jiang, 1

More information

using deep learning models to understand visual cortex

using deep learning models to understand visual cortex using deep learning models to understand visual cortex 11-785 Introduction to Deep Learning Fall 2017 Michael Tarr Department of Psychology Center for the Neural Basis of Cognition this lecture A bit out

More information

Stimulus-driven population activity patterns in macaque primary visual cortex

Stimulus-driven population activity patterns in macaque primary visual cortex Stimulus-driven population activity patterns in macaque primary visual cortex Benjamin R. Cowley a,, Matthew A. Smith b, Adam Kohn c, Byron M. Yu d, a Machine Learning Department, Carnegie Mellon University,

More information

Describe what is meant by a placebo Contrast the double-blind procedure with the single-blind procedure Review the structure for organizing a memo

Describe what is meant by a placebo Contrast the double-blind procedure with the single-blind procedure Review the structure for organizing a memo Please note the page numbers listed for the Lind book may vary by a page or two depending on which version of the textbook you have. Readings: Lind 1 11 (with emphasis on chapters 10, 11) Please note chapter

More information

Research Methods in Forest Sciences: Learning Diary. Yoko Lu December Research process

Research Methods in Forest Sciences: Learning Diary. Yoko Lu December Research process Research Methods in Forest Sciences: Learning Diary Yoko Lu 285122 9 December 2016 1. Research process It is important to pursue and apply knowledge and understand the world under both natural and social

More information